Está en la página 1de 29

Set Theory

First published Thu Jul 11, 2002

Set Theory is the mathematical science of the infinite. It studies properties of sets, abstract objects that pervade the whole of modern mathematics. The language of set theory, in its simplicity, is sufficiently universal to formalize all mathematical concepts and thus set theory, along with Predicate Calculus, constitutes the true Foundations of Mathematics. As a mathematical theory, Set Theory possesses a rich internal structure, and its methods serve as a powerful tool for applications in many other fields of Mathematics. Set Theory, with its emphasis on consistency and independence proofs, provides a gauge for measuring the consistency strength of various mathematical statements. There are four main directions of current research in set theory, all intertwined and all aiming at the ultimate goal of the theory: to describe the structure of the mathematical universe. They are: inner models, independence proofs, large cardinals, and descriptive set theory. See the relevant sections in what follows.

1. The Essence of Set Theory 2. Origins of Set Theory 3. The Continuum Hypothesis 4. Axiomatic Set Theory 5. The Axiom of Choice 6. Inner Models 7. Independence Proofs 8. Large Cardinals 9. Descriptive Set Theory Bibliography Academic Tools Other Internet Resources Related Entries

1. The Essence of Set Theory


The objects of study of Set Theory are sets. As sets are fundamental objects that can be used to define all other concepts in mathematics, they are not defined in terms of more fundamental concepts. Rather, sets are introduced either informally, and are understood as something self-evident, or, as is now standard in modern mathematics, axiomatically, and their properties are postulated by the appropriate formal axioms.

The language of set theory is based on a single fundamental relation, called membership. We say that A is a member of B (in symbols A B), or that the set B contains A as its element. The understanding is that a set is determined by its elements; in other words, two sets are deemed equal if they have exactly the same elements. In practice, one considers sets of numbers, sets of points, sets of functions, sets of some other sets and so on. In theory, it is not necessary to distinguish between objects that are members and objects that contain members the only objects one needs for the theory are sets. See the supplement Basic Set Theory for further discussion. Using the membership relation one can derive other concepts usually associated with sets, such as unions and intersections of sets. For example, a set C is the union of two sets A and B if its members are exactly those objects that are either members of Aor members of B. The set C is uniquely determined, because we have specified what its elements are. There are more complicated operations on sets that can be defined in the language of set theory (i.e. using only the relation ), and we shall not concern ourselves with those. Let us mention another operation: the (unordered) pair {A,B} has as its elements exactly the setsAand B. (If it happens that A=B, then the pair has exactly one member, and is called a singleton {A}.) By combining the operations of union and pairing, one can produce from any finite list of sets the set that contains these sets as members: {A,B,C,D,...,K,L,M}. We also mention the empty set, the set that has no elements. (The empty set is uniquely determined by this property, as it is the only set that has no elements this is a consequence of the understanding that sets are determined by their elements.) When dealing with sets informally, such operations on sets are self-evident; with the axiomatic approach, it is postulated that such operations can be applied: for instance, one postulates that for any sets A and B, the set {A,B} exists. In order to endow set theory with sufficient expressive power one needs to postulate more general construction principles than those alluded to above. The guiding principle is that any objects that can be singled out by means of the language can be collected into a set. For instance, it is desirable to have the set of all integers that are divisible by number 3, the set of all straight lines in the Euclidean plane that are parallel to a given line, the set of all continuous real functions of two real variables etc. Thus one is tempted to postulate that given any property P, there exists a set whose members are exactly all the sets that have property P. As we shall see below, such an

assumption is logically inconsistent, and the accepted construction principles are somewhat weaker than such a postulate. One of the basic principles of set theory is the existence of an infinite set. The concept can be formulated precisely in the language of set theory, using only the membership relation, and the definition captures the accepted meaning of infinite. See the supplement on Basic Set Theory for further discussion. Using the basic construction principles, and assuming the existence of infinite sets, one can definenumbers, including integers, real numbers and complex numbers, as well as functions, functionals, geometric and topological concepts, and all objects studied in mathematics. In this sense, set theory serves as Foundations of Mathematics. The significance of this is that all questions of provability (or unprovability) of mathematical statements can be in principle reduced to formal questions of formal derivability from the generally accepted axioms of Set Theory. While the fact that all of mathematics can be reduced to a formal system of set theory is significant, it would hardly be a justification for the study of set theory. It is the internal structure of the theory that makes it worthwhile, and it turns out that this internal structure is enormously complex and interesting. Moreover, the study of this structure leads to significant questions about the nature of the mathematical universe. The fundamental concept in the theory of infinite sets is the cardinality of a set. Two sets A and B have the same cardinality if there exists a mapping from the set A onto the set B which is one-to-one, that is, it assigns each element of A exactly one element of B. It is clear that when two sets are finite, then they have the same cardinality if and only if they have the same number of elements. One can extend the concept of the number of elements to arbitrary, even infinite, sets. It is not apparent at first that there might be infinite sets of different cardinalities, but once this becomes clear, it follows quickly that the structure so described is rich indeed.

2. Origins of Set Theory


The birth of Set Theory dates to 1873 when Georg Cantor proved the uncountability of the real line. (One could even argue that the exact birthdate is December 7, 1873, the date of Cantor's letter to Dedekind informing him of his discovery.) Until then, no one envisioned the possibility that infinities come in different sizes, and moreover, mathematicians had no use for actual

infinity. The arguments using infinity, including the Differential Calculus of Newton and Leibniz, do not require the use of infinite sets, and infinity appears only as a manner of speaking, to paraphrase Friedrich Gauss. The fact that the set of all positive integers has a proper subset, like the set of squares {1, 4, 9, 16, 25,...} of the same cardinality (using modern terminology) was considered somewhat paradoxical (this had been discussed at length by Galileo among others). Such apparent paradoxes prevented Bernhard Bolzano in 1840s from developing set theory, even though some of his ideas are precursors of Cantor's work. (It should be mentioned that Bolzano, an accomplished mathematician himself, coined the word Menge (= set) that Cantor used for objects of his theory.) Motivation for Cantor's discovery of Set Theory came from his work on Fourier series (which led him to introduce ordinal numbers) and on trancendental numbers. Real numbers that are solutions of polynomial equations with integer coefficients are called algebraic, and the search was on for numbers that are not algebraic. A handful of these, called transcendental numbers, was discovered around that time, and a question arose how rare such numbers are. What Cantor did was to settle this question in an unexpected way, showing in one fell swoop that transcendental numbers are plentiful indeed. His famous proof went as follows: Let us call an infinite set A countable, if its elements can be enumerated; in other words, arranged in a sequence indexed by positive integers: a(1), a(2), a(3), , a(n), . Cantor observed that many infinite sets of numbers are countable: the set of all integers, the set of all rational numbers, and also the set of all algebraic numbers. Then he gave his ingeneous diagonal argument that proves, by contradiction, that the set of all real numbers is not countable. A consequence of this is that there exists a multitude of transcendental numbers, even though the proof, by contradiction, does not produce a single specific example. See the supplement on Basic Set Theory for further discussion. Cantor's discovery of uncountable sets led him to the subsequent development of ordinal and cardinal numbers, with their underlying order and arithmetic, as well as to a plethora of fundamental questions that begged to be answered (such as the Continuum Hypothesis). After Cantor, mathematics has never been the same.

3. The Continuum Hypothesis

As the Continuum Hypothesis has been the most famous problem in Set Theory, let me explain what it says. The smallest infinite cardinal is the cardinality of a countable set. The set of all integers is countable, and so is the set of all rational numbers. On the other hand, the set of all real numbers is uncountable, and its cardinal is greater than the least infinite cardinal. A natural question arises: is this cardinal (the continuum) the very next cardinal. In other words, is it the case that there are no cardinals between the countable and the continuum? As Cantor was unable to find any set of real numbers whose cardinal lies strictly between the countable and the continuum, he conjectured that the continuum is the next cardinal: the Continuum Hypothesis. Cantor himself spent most of the rest of his life trying to prove the Continuum Hypothesis and many other mathematicians have tried too. One of these was David Hilbert, the leading mathematician of the last decades of the 19th century. At the World Congress of Mathematicians in Paris in 1900 Hilbert presented a list of major unsolved problems of the time, and the Continuum Hypothesis was the very first problem on Hilbert's list. Despite the effort of a number of mathematicians, the problem remained unsolved until 1963, and it can be argued that in some sense the problem is still unsolved. See Section 7 on Independence Proofs.

4. Axiomatic Set Theory


In the years following Cantor's discoveries, development of Set Theory proceeded with no particular concern about how exactly sets should be defined. Cantor's informal definition was sufficient for proofs in the new theory, and the understanding was that the theory can be formalized by rephrasing the informal definition as a system of axioms. In the early 1900s it became clear that one has to state precisely what basic assumptions are made in Set Theory; in other words, the need has arisen to axiomatize Set Theory. This was done by Ernst Zermelo, and the immediate reasons for his axioms were twofold. The first one was the discovery of a paradox in Set Theory. This paradox is referred to as Russell's Paradox. Consider the set S of all sets that are not an element of itself. If one accepts the principle that all such sets can be collected into a set, then S should be a set. It is easy to see however that this leads to a contradiction (is the set S an element of itself?) Russell's Paradox can be avoided by a careful choice of construction principles, so that one has the expressive power needed for usual mathematical arguments while preventing the existence of paradoxical sets. See the supplement on Zermelo-Fraenkel Set Theory

for further discussion. The price one has to pay for avoiding inconsistency is that some sets do not exist. For instance, there exists no universal set (the set of all sets), no set of all cardinal numbers, etc. The other reason for axioms was more subtle. In the course of development of Cantor's theory of cardinal and ordinal numbers a question was raised whether every set can be provided with a certain structure, called well-ordering of the set. Zermelo proved that indeed every set can be well-ordered, but only after he introduced a new axiom that did not seem to follow from the other, more self-evident, principles. His Axiom of Choice has become a standard tool of modern mathematics, but not without numerous objections of some mathematicians and discussions in both mathematical and philosophical literature. The history of the Axiom of Choice bears strong resemblance to that of the other notorious axiom, Euclid's Fifth Postulate.

5. The Axiom of Choice


The Axiom of Choice states that for every set of mutually disjoint nonempty sets there exists a set that has exactly one member common with each of these sets. For instance, let S be a set whose members are mutually disjoint finite sets of real numbers. We can choose in each X S the smallest number, and thus form a set that has exactly one member in common with each X S. What is not self-evident is whether we can make a choice every time, simultaneously for infinitely many sets X, regardless what these abstract sets are. The Axiom of Choice, which postulates the existence of a certain set (the choice set) without giving specific instructions how to construct such a set, is of a different nature than the other axioms, which all formulate certain construction principles for sets. It was this nonconstructive nature of the Axiom of Choice that fed the controversy for years to come. An interesting application of the Axiom of Choice is the Banach-Tarski Paradox that states that the unit ball can be partitioned into a finite number of disjoint sets which then can be rearranged to form two unit balls. This is of course a paradox only when we insist on visualizing abstract sets as something that exists in the physical world. The sets used in the Banach-Tarski Paradox are not physical objects, even though they do exist in the sense that their existence is proved from the axioms of mathematics (including the Axiom of Choice). The legitimate question is whether the Axiom of Choice is consistent, that is whether it cannot be refuted from the other axioms. (Notice the similarity with the non Euclidean geometry.) This question was answered by Gdel, and

eventually the role of the Axiom of Choice has been completely clarified. See Section 7 on Independence Proofs.

6. Inner Models
In the 1930s, Gdel stunned the mathematical world by discovering that mathematics is incomplete. His Incompleteness Theorem states that every axiomatic system that purports to describe mathematics as we know it must be incomplete, in the sense that one can find a true statement expressible in the system that cannot be formally proved from the axioms. In view of this result one must consider the possibility that a mathematical conjecture that resists a proof might be an example of such an unprovable statement, and Gdel immediately embarked on the project of showing that the Continuum Hypothesis might be undecidable in the axiomatic set theory. Several years after proving the Incompleteness Theorem, Gdel proved another groundbreaking result: he showed that both the Axiom of Choice and the Continuum Hypothesis are consistent with the axioms of set theory, that is that neither can be refuted by using those axioms. This he achieved by discovering a model of set theory in which both the Axiom of Choice and the Continuum Hypothesis are true. Gdel's model L of constructible sets has since served as a blueprint for building so-called inner models. These models form a hierarchy, corresponding to the hierarchy of large cardinals (see Section 8), and provide a glimpse into the as yet hidden structure of the mathematical universe. The advances in Inner Model Theory that have been made in the recent past owe much to the work of Ronald Jensen who introduced the study of the fine structure of constructible sets.

7. Independence Proofs
In 1963, Paul Cohen proved independence of the Axiom of Choice and of the Continuum Hypothesis. This he did by applying the method of forcing that he invented and constructing first a model of set theory (with the axiom of choice) in which the Continuum Hypothesis fails, and then a model of set theory in which the Axiom of Choice fails. Together with Gdel's models, these models show that the Axiom of Choice can neither be proved nor refuted from the other axioms, and that the Continuum Hypothesis can neither be proved nor refuted from the axioms of set theory (including the Axiom of Choice).

Cohen's method proved extremely fruitful and led first to the solution of a number of outstanding problems (Suslin's Problem, the Lebesgue measurability Problem, Borel's Conjecture, Kaplansky's Conjecture, Whitehead's Problem and so on) and soon has become one of the cornerstones of modern set theory. The technique of forcing has to date been applied by hundreds of authors of numerous articles and has enormously advanced our knowledge of Foundations of Mathematics. Along with the theory of large cardinals it is used to gauge the consistency strength of mathematical statements.

8. Large Cardinals
In 1930, while working on the Measure Problem, Stanislaw Ulam discovered an important phenomenon: Assuming that a certain mathematical statement about small sets (such as sets of real numbers) is true, one can prove the existence of sets of enormous size (inaccessible). This phenomenon has become more apparent after Dana Scott's celebrated result (1961) that measurable cardinals do not exist in L. Suddenly, large cardinals such as inaccessible, measurable, supercompact etc. have become the main focus of attention of set theorists. What emerged is a hierarchy of properties of infinite sets, the Large Cardinal Theory, that appears to be the basis for the structure of the set theoretical universe. Large cardinal axioms (also referred to as axioms of strong infinity) form a hierarchy whereby a stronger axiom not only implies a weaker axiom but also proves its consistency. To date there are scores of examples of mathematical statements whose consistency strength can be precisely calculated in terms of the hierarchy of large cardinals. (For instance, a negative solution of the Singular Cardinal Problem corresponds to a large cardinal axiom between measurabily and supercompactness.) Since the pioneering work of Ronald Jensen, Large Cardinal Theory has been closely tied with Inner Model Theory. It turns out that for each large cardinal axiom at lower levels of the hierarchy one can find an appropriate inner model. These inner models shed additional light on the structure of the universe by employing methods of Descriptive Set Theory.

9. Descriptive Set Theory


Descriptive Set Theory traces its origins to the theory of integration by Henri Lebesgue at the beginning of 20th century. Investigations into Borel sets of real numbers led to the theory of projective sets, and more generally, the theory of definable sets of real numbers. Following Gdel's work, it became apparent that many natural questions in Descriptive Set Theory are undecidable in axiomatic set theory. This was further confirmed by a

proliferation of independence results following Cohen's invention of the forcing method. Modern Descriptive Set Theory revolves mostly around the powerful method using infinite games. The branch of Descriptive Set Theory known as Determinateness, developed by D. A. Martin, Robert Solovay and others, brought together methods of, among others, Recursion Theory and Large Cardinal Theory and has been very successful in describing the structure of definable sets. More importantly, Descriptive Set Theory provides strong evidence for the large cardinal axioms.

Bibliography

Cantor, G., 1932, Gesammelte Abhandlungen, Berlin: Springer-Verlag. Ulam, S., 1930, Zur Masstheorie in der allgemeinen Mengenlehre, Fund. Math., 16: 140150. Gdel, K., 1938, The consistency of the axiom of choice and the generalized continuum hypothesis, Proceedings of the National Academy of Sciences (U.S.A), 24: 556557. Scott, D., 1961, Measurable cardinals and constructible sets, Bull. Acad. Pol. Sci., 9: 521524. Cohen, P., 1966, Set theory and the continuum hypothesis, New York: Benjamin. Jensen, R., 1972, The fine structure of the constructible hierarchy, Ann. Math. Logic, 4: 229308. Martin, D. and Steel, J., 1989, A proof of projective determinacy, J. Amer. Math. Soc., 2: 71125. Hrbacek, K. and Jech, T., 1999, Introduction to Set Theory, New York: Marcel Dekker, Inc.

Academic Tools
How to cite this entry. Preview the PDF version of this entry at the Friends of the SEP Society. Look up this entry topic at the Indiana Philosophy Ontology Project (InPhO). Enhanced bibliography for this entry at PhilPapers, with links to its database.

Other Internet Resources


Set Theory, maintained by Jean Larson (Mathematics, University of Florida) Articles by J.J. O'Connor and E.F. Robertson, in The MacTutor History of Mathematics archive, (Mathematics, University of St. Andrews):

A History of Set Theory o Georg Ferdinand Ludwig Philipp Cantor o Paul Joseph Cohen o Kurt Gdel o Ernst Friedrich Ferdinand Zermelo o Bernard Placidus Johann Nepomuk Bolzano A Homepage for the Axiom of Choice, maintained by Eric Schechter (Mathematics, Vanderbilt University) Gdel's Incompleteness Theorem, maintained by Dale Myers (Mathematics, University of Hawaii)
o

[Please contact the author with suggestions.]

Related Entries
Frege, Gottlob: logic, theorem, and foundations for arithmetic | logic: classical | proof theory | Russell's paradox | set theory: early development

The Early Development of Set Theory


First published Tue Apr 10, 2007; substantive revision Wed Jul 6, 2011

Set theory is one of the greatest achievements of modern mathematics. Basically all mathematical concepts, methods, and results admit of representation within axiomatic set theory. Thus set theory has served quite a unique role by systematizing modern mathematics, and approaching in a unified form all basic questions about admissible mathematical arguments including the thorny question of existence principles. This entry covers in outline the convoluted process by which set theory came into being, covering roughly the years 1850 to 1930. In 1910, Hilbert wrote that set theory is that mathematical discipline which today occupies an outstanding role in our science, and radiates [ausstrmt] its powerful influence into all branches of mathematics. This already suggests that, in order to discuss the early history, it is necessary to distinguish two aspects of set theory: its role as a fundamental language and repository of the basic principles of modern mathematics; and its role as an independent branch of mathematics, classified (today) as a branch of mathematical logic. Both aspects are considered here. The first section examines the origins and emergence of set theoretic mathematics around 1870; this is followed by a discussion of the period of

expansion and consolidation of the theory up to 1900. Section 3 provides a look at the critical period in the decade 1897 to 1908, and Section 4 deals with the time from Zermelo to Gdel (from theory to metatheory), with special attention to the often overlooked, but crucial, descriptive set theory.

1. Emergence 2. Consolidation 3. Critical Period 4. From Zermelo to Gdel Bibliography Academic Tools Other Internet Resources Related Entries

1. Emergence
The concept of a set appears deceivingly simple, at least to the trained mathematician, and to such an extent that it becomes difficult to judge and appreciate correctly the contributions of the pioneers. What cost them much effort to produce, and took the mathematical community considerable time to accept, may seem to us rather self-explanatory or even trivial. Three historical misconceptions that are widespread in the literature should be noted at the outset: (1) It is not the case that actual infinity was universally rejected before Cantor. (2) Set-theoretic views did not arise exclusively from analysis, but emerged also in algebra, number theory, and geometry. (3) In fact, the rise of set-theoretic mathematics preceded Cantor's crucial contributions. All of these points shall become clear in what follows. The notion of a collection is as old as counting, and logical ideas about classes have existed since at least the tree of Porphyry (3rd century C.E.). Thus it becomes difficult to sort out the origins of the concept of set. But sets are neither collections in the everyday sense of this word, nor classes in the sense of logicians before the mid-19th century. Ernst Zermelo, a crucial figure in our story, said that the theory had historically been created by Cantor and Dedekind. This suggests a good pragmatic criterion: one should start from authors who have significantly influenced the conceptions of Cantor, Dedekind, and Zermelo. For the most part, this is the criterion adopted here. Nevertheless, as every rule calls for an exception, the case of Bolzano is important and instructive, even though Bolzano did not significantly influence later writers.

In 19th century German-speaking areas, there were some intellectual tendencies that promoted the acceptance of the actual infinite (e.g., a revival of Leibniz's thought). In spite of Gauss's warning that the infinite can only be a manner of speaking, some minor figures and three major ones (Bolzano, Riemann, Dedekind) preceded Cantor in fully accepting the actual infinite in mathematics. Those three authors were active in promoting a set-theoretic formulation of mathematical ideas, with Dedekind's contribution in a good number of classic writings (1871, 1872, 1876/77, 1888) being of central importance. Chronologically, Bernard Bolzano was the first, but he exerted almost no influence. The high quality of his work in logic and the foundations of mathematics is well known. A book entitled Paradoxien des Unendlichen was posthumously published in 1851. Here Bolzano argued in detail that a host of paradoxes surrounding infinity are logically harmless, and mounted a forceful defence of actual infinity. He proposed an interesting argument attempting to prove the existence of infinite sets, which bears comparison with Dedekind's later argument (1888). Although he employed complicated distinctions of different kinds of sets or classes, Bolzano recognized clearly the possibility of putting two infinite sets in one-to-one correspondence, as one can easily do, e.g., with the intervals [0, 5] and [0, 12] by the function 5y = 12x. However, Bolzano resisted the conclusion that both sets are equal with respect to the multiplicity of their parts. In all likelihood, traditional ideas of measurement were still too powerful in his way of thinking, and thus he missed the discovery of the concept of cardinality. The case of Bolzano suggests that a liberation from metric concepts (which came with the development of theories of projective geometry and especially of topology) was to have a crucial role in making possible the abstract viewpoint of set theory. Bernhard Riemann proposed visionary ideas about topology, and about basing all of mathematics on the notion of set or manifold in the sense of class (Mannigfaltigkeit), in his celebrated inaugural lecture On the Hypotheses which lie at the Foundations of Geometry (1854/1868). Also characteristic of Riemann was a great emphasis on conceptual mathematics, particularly visible in his approach to complex analysis (which again went deep into topology). To give but the simplest example, Riemann was an enthusiastic follower of Dirichlet's idea that a function has to be conceived as an arbitrary injective correspondence between numerical values, be it representable by a formula or not; this meant leaving behind the times when a function was defined to be an analytic expression. Through this new style of mathematics, and through his vision of a new role

for sets and a full program for developing topology, Riemann was a crucial influence on both Dedekind and Cantor. The five-year period 18681872 saw a mushrooming of set-theoretic proposals in Germany, so much so that we could regard it as the birth of settheoretic mathematics. Riemann's geometry lecture, delivered in 1854, was published by Dedekind in 1868, jointly with Riemann's paper on trigonometric series (1854/1868b, which presented the Riemann integral). The latter was the starting point for deep work in real analysis, commencing the study of seriously discontinuous functions. The young Georg Cantor entered into this area, which led him to the study of point-sets. In 1872 Cantor introduced an operation upon point sets (see below) and soon he was ruminating about the possibility to iterate that operation to infinity and beyond: it was the first glimpse of the transfinite realm. Meanwhile, another major development had been put forward by Richard Dedekind in 1871. In the context of his work on algebraic number theory, Dedekind introduced an essentially set-theoretic viewpoint, defining fields and ideals of algebraic numbers. These ideas were presented in a very mature form, making use of set operations and of structure-preserving mappings (see a relevant passage in Ferreirs 1999, 9293; Cantor employed Dedekind's terminology for the operations in his own work on set theory around 1880: op. cit., 204). Considering the ring of integers in a given field of algebraic numbers, Dedekind defined certain subsets called ideals and operated on these sets as new objects. This procedure was the key to his general approach to the topic. In other works, he dealt very clearly and precisely with equivalence relations, partition sets, homomorphisms, and automorphisms. Thus, many of the usual set-theoretic procedures of twentieth-century mathematics go back to his work. Several years later (in 1888), Dedekind would publish a presentation of the basic elements of set theory, making only a bit more explicit the operations on sets and mappings he had been using since 1871. The following year, Dedekind published a paper in which he provided an axiomatic analysis of the structure of the set R of real numbers. He defined it as an ordered field that is also complete (in the sense that all Dedekind-cuts on R correspond to an element in R); completeness in that sense has the Archimedean axiom as a consequence. Cantor too provided a definition of Rin 1872, employing Cauchy sequences of rational numbers, which was an elegant simplification of the definition offered by Carl Weierstrass in his lectures. The form of completeness axiom that Weierstrass preferred was Bolzano's principle that a sequence of nested closed intervals in R (a sequence

such that [am+1,bm+1] [am,bm]) contains at least one real number (or, as we would say, has a non-empty intersection). The Cantor and Dedekind definitions of the real numbers relied implicitly on set theory, and can be seen in retrospect to involve the assumption of a Power Set principle. Both took as given the set of rational numbers, and for the definition of R they relied on a certain totality of infinite sets of rational numbers (either sequences, or Dedekind cuts). With this, too, constructivistic criticism of set theory began to emerge, as Leopold Kronecker started to make objections to such infinitary procedures. Simultaneously, there began a study of the topology of R, in particular in the work of Weierstrass, Dedekind, and Cantor. The set-theoretic approach was also exploited by Dedekind in his pioneering work on algebraic number theory (and algebraic geometry in 1882), and by several authors in the fields of real analysis and complex analysis. Cantor's derived sets are of particular interest (for the context of this idea in real analysis, see e.g. Dauben 1979, Hallett 1984, Lavine 1994, Kanamori 1996, Ferreirs 1999). Cantor took as given the conceptual sphere of the real numbers, and he considered arbitrary subsets P, which he called point sets. A real number r is called a limit point of P, when all neighbourhoods of r contain points of P. This can only happen if P is infinite. With that concept, introduced by Weierstrass, thederived set P of P can be defined, as Cantor did, to be the set of all the limit points of P. In general P may be infinite and have its own limit points. Thus one can iterate the operation and obtain further derived sets P, P P(n) It is easy to give examples of a set P that will give rise to non-empty derived sets P(n) for all finite n. (A rather trivial example is P = Q[0,1], the set of rational numbers in the unit interval; in this case P = [0,1] = P.) Thus one can define P() as the intersection of all P(n)for finite n. This was Cantor's first encounter with transfinite iterations. Then, in late 1873, came a surprising discovery that fully opened the realm of the transfinite. In correspondence with Dedekind (see Ewald 1996, vol. 2), Cantor asked the question whether the infinite sets N of the natural numbers and R of real numbers can be placed in one-to-one correspondence. In reply, Dedekind offered a surprising proof that the set A of all algebraic numbers is denumerable (i.e., there is a one-to-one correspondence with N). A few days later, Cantor was able to prove that the assumption that R is denumerable leads to a contradiction. To this end, he employed the Bolzano-Weierstrass principle of completeness mentioned above. Thus he had shown that there are more elements in R than in N or Q or A, in the precise sense that the cardinality of R is strictly greater than that of N.

2. Consolidation
Set theory was beginning to become an essential ingredient of the new modern approach to mathematics. But this viewpoint was contested, and its consolidation took a rather long time. Dedekind's algebraic style only began to find followers in the 1890s; David Hilbert was among them. The soil was better prepared for the modern theories of real functions: Italian, German, French and British mathematicians contributed during the 1880s. And the new foundational views were taken up by Peano and his followers, by Frege to some extent, by Hilbert in the 1890s, and later by Russell. Meanwhile, Cantor spent the years 1878 to 1885 publishing key works that helped turn set theory into an autonomous branch of mathematics. Let's write A B in order to express that the two sets A, B can be put in one-to-one correspondence (have the same cardinality). After proving that the irrational numbers can be put in one-to-one correspondence with R, and, surprisingly, that also Rn R, Cantor conjectured in 1878 that any subset of R would be either denumerable ( N) or R. This is the first and weakest form of the celebrated Continuum Hypothesis. During the following years, Cantor explored the world of point sets, introducing several important topological ideas (e.g., perfect set, closed set, isolated set), and arrived at results such as the Cantor-Bendixson theorem. A point set P is closed iff its derived set P P, and perfect iff P = P. The Cantor-Bendixson theorem then states that a closed point set can be decomposed into two subsets R and S, such that R is denumerable and S is perfect (indeed, S is the athderived set of P, for a countable ordinal a). Because of this, closed sets are said to have the perfect set property. Furthermore, Cantor was able to prove, in (1884), that perfect sets have the power of the continuum. Both results implied that the Continuum Hypothesis is valid for all closed point sets. Many years later, in 1916, Pavel Aleksandrov and Felix Hausdorff were able to show that the broader class of Borel sets have the perfect set property too. His work on points sets led Cantor, in 1882, to conceive of the transfinite numbers (see Ferreirs 1999, 267ff). This was a turning point in his research, for from then onward he studied abstract set theory independently of more specific questions having to do with point sets and their topology (until the mid-1880s, these questions had been prominent in his agenda). Subsequently, Cantor focused on the transfinite cardinal and ordinal numbers, and on general order types, independently of the topological properties of R.

The transfinite ordinals were introduced as new numbers in an important mathematico-philosophical paper of 1883,Grundlagen einer allgemeinen Mannichfaltgkeitslehre (notice that Cantor still uses Riemann's term). Cantor defined them by means of two generating principles: the first (1) yields the successor a+1 for any given number a, while the second (2) stipulates that there is a number b which follows immediately after any given sequence of numbers without a last element. Thus, after all the finite numbers comes, by (2), the first transfinite number, (read: omega); and this is followed by +1, +2, , + = 2, , n, n +1, ,2, 2+1, , , and so on and on. Whenever a sequence without last element appears, one can go on and, so to say, jump to a higher stage by (2). The introduction of these new numbers seemed like idle speculation to most of his contemporaries, but for Cantor they served two very important functions. To this end, he classified the transfinite ordinals as follows: the first number class consisted of the finite ordinals, the set N of natural numbers; the second number class was formed by and all numbers following it (including , and many more) that have only a denumerable set of predecessors. This crucial condition was suggested by the problem of proving the Cantor-Bendixson theorem. On that basis, Cantor could establish the results that the cardinality of the second number class is greater than that of N; and that no intermediate cardinality exists. Thus, if you write card(N) = 0 (read: aleph zero), his theorems justified calling the cardinality of the second number class 1. After the second number class comes a third number class (all transfinite ordinals whose set of predecessors has cardinality 1); the cardinality of this new number class can be proved to be 2. And so on. The first function of the transfinite ordinals was, thus, to establish a well-defined scale of increasing transfinite cardinalities. (The aleph notation used above was introduced by Cantor only in 1895.) This made it possible to formulate much more precisely the problem of the continuum; Cantor's conjecture became the hypothesis that card(R) = 1. Furthermore, relying on the transfinite ordinals, Cantor was able to prove the Cantor-Bendixson theorem, rounding out the results on point sets that he had been elaborating during these crucial years. The study of the transfinite ordinals directed Cantor's attention towards ordered sets, and in particular well-ordered sets. A setS is well-ordered by a relation < iff < is a total order and every subset of S has a least element in the <-ordering. (The real numbers are not well-ordered in their usual order: just consider an open interval. Meanwhile, N is the simplest infinite well-ordered set.) Cantor argued that the transfinite ordinals truly deserve the name

of numbers, because they express the type of order of any possible wellordered set. Notice also that it was easy for Cantor to indicate how to reorder the natural numbers so as to make them correspond to the order types +1, +2, , 2, , n, , 2, , , and so on. (For instance, reordering N in the form: 2, 4, 6, , 5, 15, 25, 35, , 1, 3, 7, 9, we obtain a set that has order type 3.) Notice too that the Continuum Hypothesis, if true, would entail that the set R of real numbers can indeed be well-ordered. Cantor was so committed to this viewpoint, that he presented the further hypothesis that every set can be well-ordered as a fundamental and momentous law of thought. Some years later, Hilbert called attention to both the Continuum Hypothesis and the wellordering problem as Problem 1 in his celebrated list of Mathematische Probleme (1900). Doing so was an intelligent way of emphasizing the importance of set theory for the future of mathematics, and the fruitfulness of its new methods and problems. In 1895 and 1897, Cantor published his last two articles. They were a wellorganized presentation of his results on the transfinite numbers (cardinals and ordinals) and their theory, and also on order types and well-ordered sets. However, these papers did not advance significant new ideas. Unfortunately, Cantor had doubts about a third part he had prepared, which would have discussed very important issues having to do with the problem of wellordering and the paradoxes (see below). Surprisingly, Cantor also failed to include in the 1895/97 papers a theorem which he had published some years before which is known simply as Cantor's Theorem: given any set S, there exists another set whose cardinality is greater (this is the power setp(S), as we now say Cantor used instead the set of functions f: S {0, 1}). In the same short paper (1892), Cantor presented his famous proof that R is nondenumerable by the method of diagonalisation, a method which he then extended to prove Cantor's Theorem. (A related form of argument had appeared earlier in the work of P. du Bois-Reymond.) Meanwhile, other authors were exploring the possibilities opened by set theory for the foundations of mathematics. Most important was Dedekind's contribution (1888) with a deep presentation of the theory of the natural numbers. He formulated some basic principles of set (and mapping) theory; gave axioms for the natural number system; proved that mathematical induction is conclusive and recursive definitions are flawless; developed the basic theory of arithmetic; introduced the finite cardinals; and proved that his axiom system is categorical. His system had four axioms. Given a function

defined on S, a set N S, and a distinguished element 1 N, they are as follows: () (N) N () N = o{1} () 1 (N) () the function is injective. Condition () is crucial since it ensures minimality for the set of natural numbers, which accounts for the validity of proofs by mathematical induction. N = o{1} is read: N is the chain of singleton {1} under the function , that is, the minimal closure of {1} under the function . In general, one considers the chain of a set A under an arbitrary mapping , denoted by o(A); in his booklet Dedekind developed an interesting theory of such chains, which allowed him to prove the Cantor-Bernstein theorem. The theory was later generalized by Zermelo. In the following years, Giuseppe Peano gave a more superficial (but also more famous) treatment of the natural numbers, employing the new symbolic language of logic, and Gottlob Frege elaborated his own ideas, which however fell prey to the paradoxes. An important book inspired by the set-theoretic style of thinking was Hilbert's Grundlagen der Geometrie (1899), which took the mathematics of axioms one step beyond Dedekind through a rich study of geometric systems motivated by questions concerning the independence of his axioms. Hilbert's book made clear the new axiomatic methodology that had been shaping up in connection with the novel methods of set theory, and he combined it with the axiomatic trends coming from projective geometry.

3. Critical Period
In the late nineteenth century, it was a widespread idea that pure mathematics is nothing but an elaborate form of arithmetic. Thus it was usual to talk about the arithmetisation of mathematics, and how it had brought about the highest standards of rigor. With Dedekind and Hilbert, this viewpoint led to the idea of grounding all of pure mathematics in set theory. The most difficult steps in bringing forth this viewpoint had been the establishment of a theory of the real numbers, and a set-theoretic reduction of the natural numbers. Both problems had been solved by the work of Cantor and Dedekind. But precisely when mathematicians were celebrating that full rigor had been finally attained, serious problems emerged for the foundations of set theory. First Cantor, and then Russell, discovered the paradoxes in set theory.

Cantor was led to the paradoxes by having introduced the conceptual sphere of the transfinite numbers. Each transfinite ordinal is the order type of the set of its predecessors; e.g., is the order type of {0, 1, 2, 3, }, and +2 is the order type of {0, 1, 2, 3, , , +1}. Thus, to each initial segment of the series of ordinals, there corresponds an immediately greater ordinal. Now, the whole series of all transfinite ordinals would form a well-ordered set, and to it there would correspond a new ordinal number. This is unacceptable, for this ordinal o would have to be greater than all members of the whole series, and in particular o < o. This is usually called the Burali-Forti paradox, or paradox of the ordinals (although Burali-Forti himself failed to formulate it clearly, see Moore & Garciadiego 1981). Although Cantor might have found that paradox as early as 1883, immediately after introducing the transfinite ordinals (for arguments in favour of this idea see Purkert & Ilgauds 1987 and Tait 2000), the evidence indicates clearly that it was not until 1896/97 that he found this paradoxical argument and realized its implications. By this time, he was also able to employ Cantor's Theorem to yield the Cantor paradox, or paradox of the alephs: if there existed a set of all cardinal numbers (alephs), Cantor's Theorem applied to it would give a new aleph , such that < . The great set theorist realized perfectly well that these paradoxes were a fatal blow to the logical approaches to sets favoured by Frege and Dedekind. Cantor emphasized that his views were in diametrical opposition to Dedekind's, and in particular to his nave assumption that all well-defined collections, or systems, are also consistent systems (see the letter to Hilbert, Nov. 15, 1899, in Purkert & Ilgauds 1987, 154). (Contrary to what has often been claimed, Cantor's ambiguous definition of set in his paper of 1895 was intended to be diametrically opposite to the logicists understanding of sets often called nave set theory, or can more properly be called the dichotomy conception of sets, following a suggestion of Gdel.) Cantor thought he could solve the problem of the paradoxes by distinguishing between consistent multiplicities or sets, and inconsistent multiplicities. But, in the absence of explicit criteria for the distinction, this was simply a verbal answer to the problem. Being aware of deficiencies in his new ideas, Cantor never published a last paper he had been preparing, in which he planned to discuss the paradoxes and the problem of well-ordering (we know quite well the contents of this unpublished paper, as Cantor discussed it in correspondence with Dedekind and Hilbert; see the 1899 letters to Dedekind in Cantor 1932, Ewald 1996, vol. II, or Cantor 2005). Cantor presented an argument that relied on the Burali-Forti paradox of the ordinals, and aimed to prove that every set can be well-ordered. This argument was later

rediscovered by the British mathematician P. E. B. Jourdain, but it is open to criticism because it works with inconsistent multiplicities. Cantor's paradoxes convinced Hilbert and Dedekind that there were important doubts concerning the foundations of set theory. Hilbert formulated a paradox of his own (Peckhaus & Kahle 2002), and discussed the problem with mathematicians in his Gttingen circle. Ernst Zermelo was thus led to discover the paradox of the set of all sets that are not members of themselves (Rang & Thomas 1981). This was independently discovered by Bertrand Russell, who was led to it by a careful study of Cantor's Theorem, which conflicted deeply with Russell's belief in a universal set. Some time later, in June 1902, he communicated the contradiction to Gottlob Frege, who was completing his own logical foundation of arithmetic, in a wellknown letter. Frege's reaction made very clear the profound impact of this contradiction upon the logicist program. Can I always speak of a class, of the extension of a concept? And if not, how can I know the exceptions? Faced with this, I cannot see how arithmetic could be given a scientific foundation, how numbers could be conceived as logical objects (Frege 1903, 253). The publication of Volume II of Frege's Grundgesetze, and above all Russell's work The Principles of Mathematics, made the mathematical community fully aware of the existence of the set-theoretic paradoxes, of their impact and importance. There is evidence that, up to then, even Hilbert and Zermelo had not fully appreciated the damage. Notice that the Russell-Zermelo paradox operates with very basic notions negation and set membership concepts that had widely been regarded as purely logical. The set R = {x: x x} exists according to the principle of comprehension (which allows any open sentence to determine a class), but if so, R R iff R R. It is a direct contradiction to the principle favoured by Frege and Russell. It was obviously necessary to clarify the foundations of set theory, but the overall situation did not make this an easy task. The different competing viewpoints were widely divergent. Cantor had a metaphysical understanding of set theory and, although he had one of the sharpest views of the field, he could not offer a precise foundation. It was clear to him (as it had been, somewhat mysteriously, to Ernst Schrder in his Vorlesungen ber die Algebra der Logik, 1891) that one has to reject the idea of a Universal Set, favoured by Frege and Dedekind. Frege and Russell based their approach on the principle of comprehension, which was shown contradictory. Dedekind avoided that principle, but he postulated that the Absolute Universe was a set, a thing in his technical sense of Gedankending; and he coupled that assumption with full acceptance of arbitrary subsets.

This idea of admitting arbitrary subsets had been one of the deep inspirations of both Cantor and Dedekind, but none of them had thematized it. (Here, their modern understanding of analysis played a crucial but implicit background role, since they worked within the Dirichlet-Riemann tradition of arbitrary functions.) As for the now famous iterative conception there were some elements of it (particularly in Dedekind's work, with his iterative development of the number system, and his views on systems and things), but it was conspicuously absent from many of the relevant authors. Typically, e. g., Cantor did not iterate the process of set formation: he tended to consider sets of homogeneous elements, elements which were taken to belong in some conceptual sphere (either numbers, or points, or functions, or even physical particles but not intermingled). The iterative conception was first suggested by Kurt Gdel in 1933, in connection with technical work by von Neumann and Zermelo a few years earlier; Gdel would insist on the idea in his wellknown paper on Cantor's continuum problem. It came only post facto, after very substantial amounts of set theory had been developed and fully systematized. This variety of conflicting viewpoints contributed much to the overall confusion, but there was more. In addition to the paradoxes discussed above (set-theoretic paradoxes, as we say), the list of logical paradoxes included a whole array of further ones (later called semantic). Among these are paradoxes due to Russell, Richard, Knig, Berry, Grelling, etc., as well as the ancient liar paradox due to Epimenides. And the diagnoses and proposed cures for the damage were tremendously varied. Some authors, like Russell, thought it was essential to find a new logical system that could solve all the paradoxes at once. This led him into the ramified type theory that formed the basis of Principia Mathematica (3 volumes, 19101913), his joint work with Alfred Whitehead. Other authors, like Zermelo, believed that most of those paradoxes dissolved as soon as one worked within a restricted axiomatic system. They concentrated on the set-theoretic paradoxes (as we have done above), and were led to search for axiomatic systems of set theory. Even more importantly, the questions left open by Cantor and emphasized by Hilbert in his first problem of 1900 caused heated debate. At the International Congress of Mathematicians at Heidelberg, 1904, Gyula (Julius) Knig proposed a very detailed proof that the cardinality of the continuum cannot be any of Cantor's alephs. His proof was only flawed because he had relied on a result previously proven by Felix Bernstein, a student of Cantor and Hilbert. It took some months for Felix Hausdorff to identify the flaw and correct it by properly stating the special conditions under which Bernstein's result was valid (see Hausdorff 2002). Once thus corrected,

Knig's theorem became one of the very few results restricting the possible solutions of the continuum problem, implying e.g. that Card(R) is not equal to . Meanwhile, Zermelo was able to present a proof that every set can be well-ordered, using the Axiom of Choice. During the following year, prominent mathematicians in Germany, France, Italy and England discussed the Axiom of Choice and its acceptability. This started a whole era during which the Axiom of Choice was treated most carefully as a dubious hypothesis (see the monumental study by Moore 1982). And that is ironic, for, among all of the usual principles of set theory, the Axiom of Choice is the only one that explicitly enforces the existence of some arbitrary subsets. But, important as this idea had been in motivating Cantor and Dedekind, and however entangled it is with classical analysis, infinite arbitrary subsets were rejected by many other authors. Among the most influential ones in the following period, one ought to emphasize the names of Russell, Hermann Weyl, and of course Brouwer. The impressive polemics which surrounded his Well-Ordering Theorem, and the most interesting and difficult problem posed by the foundations of mathematics, led Zermelo to concentrate on axiomatic set theory. As a result of his incisive analysis, in 1908 he published his axiom system, showing how it blocked the known paradoxes and yet allowed for a masterful development of the theory of cardinals (and ordinals). This, however, is the topic of another entry.

4. From Zermelo to Gdel


In the period 19001930, the rubric set theory was still understood to include topics in topology and the theory of functions. Although Cantor, Dedekind, and Zermelo had left that stage behind to concentrate on pure set theory, for mathematicians at large this would still take a long time. Thus, at the first International Congress of Mathematicians, 1897, keynote speeches given by Hadamard and Hurwitz defended set theory on the basis of its importance for analysis. Around 1900, motivated by topics in analysis, important work was done by three French experts: Borel, Baire and Lebesgue. Their work inaugurated the development of descriptive set theory by extending Cantor's studies on definable sets of real numbers (in which he had established that the Continuum Hypothesis is valid for closed sets). They introduced the hierarchy of Borel sets, the Baire hierarchy of functions, and the concept of Lebesgue measure a crucial concept of modern analysis. Descriptive set theory (DST) is the study of certain kinds of definable sets of real numbers, which are obtained from simple kinds (like the open sets and the

closed sets) by well-understood operations like complementation or projection. The Borel setswere the first hierarchy of definable sets, introduced in the 1898 book of mile Borel; they are obtained from the open sets by iterated application of the operations of countable union and complementation. In 1905 Lebesgue studied the Borel sets in an epochal memoir, showing that their hierarchy has levels for all countable ordinals, and analyzing the Baire functions as counterparts of the Borel sets. The main aim of descriptive set theory is to find structural properties common to all such definable sets: for instance, the Borel sets were shown to have the perfect set property (if uncountable, they have a perfect subset) and thus to comply with the continuum hypothesis (CH). This result was established in 1916 by Hausdorff and by Alexandroff, working independently. Other important regularity properties studied in DST are the property of being Lebesgue measurable, and the so-called property of Baire (to differ from an open set by a so-called meager set, or set of first category). Also crucial at the time was the study of the analytic sets, namely the continuous images of Borel sets, or equivalently, the projections of Borel sets. The young Russian mathematician Mikhail Suslin found a mistake in Lebesgue's 1905 memoir when he realized that the projection of a Borel set is not Borel in general. However, he was able to establish that the analytic sets, too, possess the perfect set property and thus verify CH. By 1923 Nikolai Lusin and Waclaw Sierpiski were studying the co-analytic sets, and this was to lead them to a new hierarchy of projective sets, which starts with the analytic sets (11), their complements (co-analytic, 11 sets), the projections of these last (12 sets), their complements (12 sets), and so on. During the 1920s much work was done on these new types of sets, mainly by Polish mathematicians around Sierpiski and by the Russian school of Lusin and his students. A crucial result obtained by Sierpiski was that every 12 set is the union of 1 Borel sets (the same holds for 11 sets), but this kind of traditional research on the topic would stagnate after around 1940. Soon Lusin, Sierpiski and their colleagues were finding extreme difficulties in their work. Lusin was so much in despair that, in a paper of 1925, he came to the totally unexpected conclusion that one does not know and one will never know whether the projective sets have the desired regularity properties (quoted in Kanamori [1993], 250). Such comments are highly interesting in the light of later developments, which have led to hypotheses that solve all the relevant questions (Projective Determinacy). They underscore the difficult methodological and philosophical issues raised by these more recent hypotheses, namely the problem concerning the kind of evidence that backs them.

Lusin summarized the state of the art in his 1930 book Leons sur les ensembles analytiques (Paris, Gauthier-Villars), which was to be a key reference for years to come. Since this work, it has become customary to present results in DST for the Baire space of infinite sequences of natural numbers, which in effect had been introduced by Ren Baire in his dissertation of 1899 (anticipated by Dedekind in 1891). Baire space is endowed with a certain topology that makes it homeomorphic to the set of the irrational numbers, and it is regarded by experts to be perhaps the most fundamental object of study of set theory next to the set of natural numbers (Moschovakis 1994). This stream of work on DST must be counted among the most important contributions made by set theory to analysis and topology. But what had begun as an attempt to prove the Continuum Hypothesis could not reach this goal. Soon it was shown using the Axiom of Choice that there are nonLebesgue measurable sets of reals (Vitali, 1905), and also uncountable sets of reals with no perfect subset (Bernstein, 1908). Such results made clear the impossibility of reaching the goal of CH by concentrating on definable and well-behaved sets of reals. Also, with Gdel's work around 1940 (and also with forcing in the 1960s) it became clear why the research of the 1920s and 30s had stagnated: the fundamental new independence results showed that the theorems established by Suslin (perfect set property for analytic sets), Sierpinski (12 sets as unions of 1 Borel sets) and a few others were the best possible results on the basis of axiom system ZFC. This is important philosophically: already an exploration of the world of sets definable from the open (or closed) sets by complement, countable union, and projection had sufficed to reach the limits of the ZFC system. Hence the need for new axioms, that Gdel emphasized after World War II. Let us now turn to Cantor's other main legacy, the study of transfinite numbers. By 1908 Hausdorff was working on uncountable order types and introduced the Generalized Continuum Hypothesis (2a = a+1). He was also the first to consider the possibility of an exorbitant cardinal, namely a weakly inaccessible, i.e., a regular cardinal that is not a successor (a cardinal is called regular if decomposing into a sum of smaller cardinals requires many such numbers). Few years later, in the early 1910s, Paul Mahlo was studying hierarchies of such large cardinals in work that pioneered what was to become a central area of set theory; he obtained a succession of inaccessible cardinals by employing a certain operation that involves the notion of a stationary subset; they are called Mahlo cardinals. But the study of

large cardinals developed slowly. Meanwhile, Hausdorff's textbook Grundzge der Mengenlehre (1914; see his 2002) introduced two generations of mathematicians into set theory and general topology. The next crucial steps into the very high infinite were done in 1930. The notion of strongly inaccessible cardinals was then isolated by Sierpinski & Tarski, and by Zermelo. A strong inaccessible is a regular cardinal such that 2x is less than whenever x < . While weak inaccessibles merely involve closure under the successor operation, strong inaccessibles involve a much stronger notion of closure under the powerset operation. That same year, in a path-breaking paper on models of ZFC, Zermelo established a link between the uncountable (strongly) inaccessible cardinals and certain natural models of ZFC (in which work he assumed, so to say, that the powerset operation is fully determined). In that same year, Stanislaw Ulam was led by considerations coming out of analysis (measure theory) to a concept that was to become central: measurable cardinals. It turned out that such cardinals, defined by a measure-theoretic property, had to be (strongly) inaccessible. Indeed, many years later it would be established (by Hanf, working upon Tarski's earlier work) that the first inaccessible cardinal is not measurable, showing that these new cardinals were even more exorbitant. As one can see, the Polish school led by Sierpiski had a very central role in the development of set theory between the Wars. Measurable cardinals came to special prominence in the late 1960s when it became clear that the existence of a measurable cardinal contradicts Gdel's axiom of constructibility (V = L in the class notation). This again vindicated Gdel's convictions, expressed in what is sometimes called Gdel's program for new axioms. Set-theoretic mathematics continued its development into the powerful axiomatic and structural approach that was to dominate much of the 20th century. To give just a couple of examples, Hilbert's early axiomatic work (e.g. in his arch-famousFoundations of Geometry) was deeply set-theoretic; Ernst Steinitz published in 1910 his research on abstract field theory, making essential use of the Axiom of Choice; and around the same time the study of function spaces began with work by Hilbert, Maurice Frchet, and others. During the 1920s and 30s, the first specialized mathematics journal, Fundamenta Mathematicae, was devoted to set theory as then understood (centrally including topology and function theory). In those decades structural algebra came of age, abstract topology was gradually becoming an independent branch of study, and the study of set theory initiated its metatheoretic turn. Ever since, set theory has generally been identified

with the branch of mathematical logic that studies transfinite sets, originating in Cantor's result that R has a greater cardinality than N. But, as the foregoing discussion shows, set theory was both effect and cause of the rise of modern mathematics: the traces of this origin are indelibly stamped on its axiomatic structure.

Bibliography

Cantor, Georg, 1883. Grundlagen einer allgemeinen Mannigfaltigkeitslehre, Leipzig: B. G. Teubner. In Cantor 1932, 165208. English in Ewald 1996, vol. 2. Cantor, Georg, 1884. ber unendliche, lineare Punktmannichfaltigkeiten, 6, Mathematische Annalen, 23: 45388. In Cantor (1932). 210244. Cantor, Georg, 1892. ber eine elementare Frage der Mannigfaltigkeitslehre, Jahresbericht der Deutschen Mathematiker Vereinigung, 1: 7578. In Cantor (1932), 278280. English trans. in [Ewald 1996], vol.2. Cantor, Georg, 1895/97. Beitrge zur Begrndung der transfiniten Mengenlehre, in Cantor 1932, 282351. English trans. in Cantor, Contributions to the founding of the theory of transfinite numbers, New York: Dover, 1955. Cantor, Georg, 1932. Gesammelte Abhandlungen mathematischen und philosophischen Inhalts, E. Zermelo (ed.), Berlin: Springer. Reprint Hildesheim: Olms, 1966. Cavaills, Jean, 1962. Philosophie mathmatique, Paris: Hermann, 1962. Dauben, Joseph, 1979. Georg Cantor. His Mathematics and Philosophy of the Infinite, Cambridge, MA: Harvard University Press. Dedekind, Richard, 1871. ber die Komposition der binren quadratischen Formen, Supplement X to G. L. Dirichlet & R. Dedekind, Vorlesungen ber Zahlentheorie, Braunschweig: Vieweg. [Later editions as Supplement XI, of which the fourth is reprinted in New York: Chelsea, 1968.] Partial reprint in Dedekind (1930/32), vol.3, 223261. Dedekind, Richard, 1872. Stetigkeit und irrationale Zahlen, Braunschweig: Vieweg. In Dedekind (1930/32), vol.3, 315334. English trans. in [Ewald 1996], vol.2. Dedekind, Richard, 1876/77. Sur la thorie des nombres entiers algbriques, Bulletin des Sciences mathmatiques et astronomiques, 1st series, XI (1876), 278293; 2nd series, I (1877), 1741, 6992, 144164, 207248. Separate edition, Paris: Gauthier-Villars, 1977. English trans. by J. Stillwell: Theory of algebraic integers, Cambridge: Cambridge University Press, 2004.

Dedekind, Richard, 1888. Was sind und was sollen die Zahlen?, Braunschweig: Vieweg. In Dedekind 1930/32, vol. 3. English in Ewald 1996, vol. 2. Dedekind, Richard, 1930/32. Gesammelte mathematische Werke, R. Fricke, E. Noether & . Ore (eds.), Braunschweig: Vieweg, 3 vols. Reprint New York: Chelsea, 1969. Ewald, William B., 1996. From Kant to Hilbert: A source book in the foundations of mathematics, 2 vols., Oxford: Oxford University Press. Ferreirs, Jos, 1999. Labyrinth of Thought. A history of set theory and its role in modern mathematics, Basel: Birkhuser. Frege, Gottlob, 1903. Grundgesetze der Arithmetik, vol. 2, Jena: Pohle. Reprint Hildesheim: Olms, 1966. Grattan-Guinness, Ivor (ed.), 1980. From the Calculus to Set Theory, 1630 1910, London: Duckworth. Hallett, Michael, 1984. Cantorian Set Theory and Limitation of Size, Oxford: Clarendon. Hausdorff, Felix, 2002. Gesammelte Werke, vol. II: Grundzge der Mengenlehre, E. Brieskorn, W. Purkert et al. (eds.), Berlin: Springer. van Heijenoort, Jean, 1967. From Frege to Gdel: A source book in mathematical logic, Cambridge, MA: Harvard University Press. Reprint as paperback, 2000. Kanamori, Akihiro, 1996. The mathematical development of set theory from Cantor to Cohen, Bulletin of Symbolic Logic, 2: 171. Kanamori, Akihiro, 1995. The emergence of descriptive set theory, Synthese, 251: 241262. . Lavine, Shaughan, 1994. Understanding the Infinite, Cambridge, MA: Harvard University Press. Maddy, Penelope, 1988. Believing the axioms, Journal of Symbolic Logic, 53 (2): 481511; 53 (3): 736764. Moore, Gregory H., 1982. Zermelo's Axiom of Choice. Its Origins, Development and Influence, Berlin: Springer. Moore, G. H. & A. Garciadiego, 1981. Burali-Forti's Paradox: A reappraisal of its origins, Historia Mathematica, 8: 31950. Moschovakis, Yiannis N., 1994. Set Theory Notes, New York: Springer. Peckhaus, Volker & R. Kahle, 2002. Hilbert's Paradox, Historia Mathematica, 29 (2): 157175. Purkert, Walter & H.J. Ilgauds, 1987. Georg Cantor 18451918, Basel: Birkhuser. Rang, Bernhard & W. Thomas, 1981. Zermelo's Discovery of the Russell Paradox, Historia Mathematica, 8: 1522. Riemann, Bernhard, 1854/1868. ber die Hypothesen, welche der Geometrie zu Grunde liegen (Habilitationsvotrag),Abhandlungen der Kniglichen

Gesellschaft der Wissenschaften zu Gttingen, 13 (1868): 133-152. In Riemann (1892), 272287. English translation by Clifford, reprinted in [Ewald 1996], vol. 2. Riemann, Bernhard, 1854/1868b. ber die Darstellbarkeit einer Function durch eine trigonometrische Reihe, (Habilitationsschrift), Abhandlungen der Kniglichen Gesellschaft der Wissenschaften zu Gttingen, 13 (1868): 87 132. In Riemann (1892), 227265. Riemann, Bernhard, 1892. Gesammelte mathematische Werke und wissenschaftlicher Nachlass, H. Weber and R. Dedekind (eds.), Leipzig, Teubner. Reprinted (together with the Nachtrge, M. Noether and W. Wirtinger (eds.), New York: Dover, 1953. Russell, Bertrand, 1903. The Principles of Mathematics, Cambridge, University Press. Reprint of the 2nd edn. (1937): London: Allen & Unwin, 1948. Sierpiski, Wacaw, 197476. Oeuvres choisies, S. Hartman, et al. (eds.), Volumes 23; Warszawa, Editions scientifiques de Pologne. Tait, William, 2000. Cantor's Grundlagen and the Paradoxes of Set Theory, W. Tait, The Provenance of Pure Reason, Oxford: Oxford University Press, 2005, pp. 252275. Zermelo, Ernst, 1904. Beweis, dass jede Menge wohlgeordnet werden kann, Mathematische Annalen, 59: 514516. English trans. in van Heijenoort 1967. Zermelo, Ernst, 1908. Untersuchungen ber die Grundlagen der Mengenlehre, Mathematische Annalen, 65: 261281. English trans. in van Heijenoort 1967.

Academic Tools
How to cite this entry. Preview the PDF version of this entry at the Friends of the SEP Society. Look up this entry topic at the Indiana Philosophy Ontology Project (InPhO). Enhanced bibliography for this entry at PhilPapers, with links to its database.

Other Internet Resources

A History of Set Theory, by J.J. O'Connor and E.F. Robertson, in The MacTutor History of Mathematics archive. Note that their reconstruction conflicts at some points with the one provided here. Godel's Program (PowerPoint), an interesting talk by John R. Steel (Mathematics, U.C./Berkeley). A Home Page for the Axiom of Choice, maintained by Eric Schechter (Mathematics, Vanderbilt University).

Related Entries
category theory | continuity and infinitesimals | Dedekind, Richard: contributions to the foundations of mathematics | Frege, Gottlob | Hilbert, David: program in the foundations of mathematics | paradoxes: and contemporary logic | Russell, Bertrand |Russell's paradox | set theory | set theory: continuum hypothesis | set theory: Zermelo's axiomatization of | Whitehead, Alfred North

También podría gustarte